Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Ultra-efficient second harmonic generation via mode phase matching in integrated lithium niobate racetrack resonators

Open Access Open Access

Abstract

High-efficiency second harmonic generation (SHG) relying solely on intermodal dispersion engineering remains a challenge. Here, we realize highly efficient SHG using a double-waveguide coupled racetrack microring resonator on X-cut lithium niobate on insulator (LNOI), where both pump and second harmonic (SH) approach critical coupling. Through precise temperature tuning, simultaneous pump and SH resonance is attained in the resonator, dramatically enhancing SHG efficiency. With low pump power, a normalized conversion efficiency of 9972%/W is achieved. Moreover, the resonator provides a 25.73 dB enhancement in SHG efficiency compared to a 4 mm straight waveguide with identical phase matching in our experiment. This work enables efficient wavelength conversion and quantum state generation on integrated X-cut LNOI platforms.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent years, lithium niobate on insulator (LNOI) [14] has emerged as one of the most promising material platforms for integrated optics. Owing to its strong second-order nonlinearity (d33 = 27 pm/V), relatively large refractive index contrast, low optical absorption, ultra-wide transparency spanning the ultraviolet to far-infrared, and unique ferroelectric properties, LNOI has become the foremost choice for second-order nonlinear optics. Diverse nonlinear optical processes enabled by LNOI include second harmonic generation (SHG) [514], sum- and difference-frequency generation [15,16], and parametric down conversion (PDC) [1720], which play critical roles in coherent light generation at arbitrary wavelengths and nonclassical quantum state preparation [18]. Among these, SHG, which doubles the frequency of an incident field by harnessing the tight modal confinement of micronano waveguides and resonance enhancement in high-quality microcavities, has constituted an area of particular research focus.

The SHG process necessitates fulfilling phase matching conditions to occur efficiently. Numerous phase matching approaches have been realized on lithium niobate on insulator (LNOI), including birefringent phase matching [21], modal phase matching [7,10], cyclic phase matching [22,23], periodic geometry engineering [6], quasi phasematching via periodic poling [2426]. Periodically poled lithium niobate (PPLN) currently constitutes the most effective platform for nonlinear frequency conversion. Using this approach, Lu et al. achieved the highest second harmonic generation (SHG) efficiency to date of 5,000,000%/W in a microring [17]. Cyclic phase matching typically necessitates ultrahigh-Q microdisk resonators (Q∼106) [23] fabricated by specialized chemical-mechanical polishing, which can only couple to tapered fibers without on-chip integration. Modal phase matching for SHG has witnessed remarkable advancements on the lithium niobate on insulator (LNOI) platform. However, there are no suitable microresonators to enhance the efficiency to the desired level. The isotropic in-plane optical properties of Z-cut LNOI facilitate designing phase-matched microring cavities. Luo et al. attained 1500%/W normalized SHG conversion efficiency via microcavity-enhanced modal phase matching on Z-cut LNOI [15]. Chen et al. realized 26%/W/cm2 SHG efficiency in phase-matched waveguides, demonstrating a 10-fold enhancement with microring resonators [10]. However, a single waveguide coupling to the microring cannot concurrently fulfill the critical coupling conditions for both the pump and the second harmonic (SH), limiting the attainable efficiency. Furthermore, microresonator-enhanced modal phase matching for SHG remains unexplored on X-cut LNOI, impeding integration of high-performance electro-optic modulators [27] and other devices enabled by this platform.

Here, we propose and experimentally validate enhancing SHG efficiency using a modal phase matched racetrack microring resonator (RMR) on X-cut LNOI. Independent couplers for the pump and second-harmonic enable simultaneous critical coupling. The RMR radius is tailored to preclude quality factor degradation from mode hybridization. Experimentally, the extracted loaded quality factors are $Q_{T{E_0}}^{1560nm} = 1 \times {10^5}$ and $Q_{T{E_2}}^{780nm} = 2.8 \times {10^4}$ for the interacting pump and SH modes. Precisely controlling the RMR temperature allows concurrent pump and SH resonance, further boosting conversion efficiency. At low pump power, a normalized efficiency of 9972%/W is attained. In our experiment, the microring resonator demonstrated a 25.73 dB enhancement in SHG efficiency compared to a 4 mm straight waveguide with the same mode phase matching design.

2. Design and simulation of double resonance RMR

As illustrated in Fig. 1(a), the schematic diagram depicts a double-resonance RMR on an X-cut LNOI wafer. The illustration in Fig. 1(a) shows the cross section of the waveguide, where the sidewall inclination of the lithium niobate waveguide is theta = 66°. This angle was precisely measured by scanning electron microscopy in our previous experiments, and is incorporated in subsequent simulations. The 1560 nm pump is coupled from the top waveguide into the RMR, where it experiences power enhancement and nonlinear frequency conversion facilitated by the ring resonator. The resulting 780 nm SH is out-coupled from the RMR through the bottom waveguide. The straight section of the RMR has a length of 600 µm, chosen to achieve phase matching between the TE0 mode of the pump and the TE2 mode of the SH light, which is $nef{f_{{\omega _1},TE0}} = nef{f_{{\omega _2},TE2}}$. This configuration allows efficient utilization of the maximum nonlinear coefficient d33 and provides a relatively large mode overlap factor compared to the asymmetric TE1 mode of the second harmonic light.

 figure: Fig. 1.

Fig. 1. (a) Schematic of double resonance RMR and the illustration shows the cross section of the LN waveguide. (b) Simulation of effective refractive index of pump and second harmonic light in straight waveguide. (c,d) Simulation of effective refractive index versus rotation angle in 180° curved waveguide. (e) FDTD simulation of TE2 and TM2 mode crosstalk for different bending radii. Insets show field distribution for radii of 26 µm and 30 µm.

Download Full Size | PDF

As shown in Fig. 1(b), by numerically solving the effective refractive index of intrinsic mode of the optical waveguide, it can be seen that when the waveguide width is 0.733 µm, the geometric dispersion of the waveguide can compensate the material dispersion to achieve phase matching between the two modes. The upper left illustration shows the mode field distribution of the two interacting modes. The X-cut LNOI has an anisotropic refractive index in the chip plane, with ordinary and extraordinary optical axis refractive indices of no = 2.2108, ne = 2.1373 at 1560 nm, and no = 2.2580, ne = 2.1778 at 780 nm [28]. The substrate material of the wafer is silicon dioxide (SiO2), with a material refractive index of 1.4439 at 1560 nm and 1.4537 at 780 nm [29]. When the light field is transmitted in the curved waveguide, and the refractive index along the xyz direction on the cross section are ${n_x} = {(\frac{{{{\sin }^2}\alpha }}{{n_0^2}} + \frac{{{{\cos }^2}\alpha }}{{n_e^2}})^{\textrm{ - }\frac{1}{2}}}$, ${n_y} = {n_0}$ and ${n_z} = {(\frac{{{{\cos }^2}\alpha }}{{n_0^2}} + \frac{{{{\sin }^2}\alpha }}{{n_e^2}})^{\textrm{ - }\frac{1}{2}}}$, where α is the angle between the optical axis and the transmission section of the optical field. The illustration in Fig. 1(b) shows the numerical simulation results of the effective refractive index of the curved waveguide. The horizontal axis is the angle between the wave vector and the ordinary optical axis, and the vertical axis is the effective refractive index of bend waveguide. Because of the variable refractive index of the bend waveguide, mode hybridization occurs when the effective refractive index of the two eigenmodes of the waveguide is very close, which is similar to mode hybridization caused by the change of waveguide width [30].

Figure 1(c) and (d) show the numerical simulation results of the effective refractive index of the pump and SH target transport modes and orthogonal modes to them. We see that the effective refractive index of TE0 mode and TM0 mode of the pump is far different, so there is no mode hybridization and crosstalk in the transmission process. However, the effective refractive index curves of the SH light of TE2 and TM2 mode show a cross phenomenon in the blue circle, that is, mode hybridization will occur. It is worth noting that this simulation result has no relation to the bending radius, that is, we can’t eliminate the phenomenon of mode hybridization by changing the bending radius. However, the phase difference between the two modes transmitted can be controlled by changing the radius, and this mode hybridization will cancel out when the phase difference is just enough to produce constructive interference between the two modes. Figure 1(e) shows the finite-difference time-domain (FDTD) simulation results for input TE2 mode in a 180 degree bend waveguide with different bending radius. The illustration shows the electric field distribution perpendicular to the transmission plane at radii of 26 µm and 30 µm. We see that the TE2 mode and TM2 mode produce the most constructive interference when the bending radius is 26 µm, with almost no intermode crosstalk after transmission in a 180 degree bend. This result makes the design of TE2 mode directional couplers easier, and can improve the quality factor and extinction ratio of TE2 mode transmission in the microring.

Efficient simultaneous coupling of two distinct optical frequencies poses a significant challenge in nonlinear microring resonators. Realizing maximum normalized conversion efficiency in SHG microrings necessitates meeting critical coupling conditions for both the pump and SH waves. In our proposed device (Fig. 1(a)), the pump and SH input and output are facilitated through a directional coupler (DC) at the top and an asymmetric directional coupler (ADC) at the bottom, respectively. As illustrated in Fig. 2(a), the 1560 nm pump couples to the RMR via a 180-degree bent waveguide. Capitalizing on the evanescent field disparity between the 1560 nm TE0 and 780 nm TE2 modes, FDTD simulations (Fig. 2(b)) demonstrate a coupling coefficient of 0.05 for the 1560 nm TE0 mode when the coupling gap g = 0.7 µm, satisfying critical coupling. The coupling coefficient for the 780 nm TE2 SH mode is an order of magnitude smaller, minimally compromising the quality factor.

 figure: Fig. 2.

Fig. 2. (a) Schematic diagram of the DC at the top of the RMR. (b)FDTD simulation results of the DC for 1560 nm pump and 780 nm SH. (c) Schematic diagram of the ADC at the bottom of the racetrack microring. The variation curves of ADC wide waveguide width ws2 (d) and coupling gap gz (e) with propagation length z. (f) FDTD simulation results of the ADC for 1560 nm pump and 780 nm SH.

Download Full Size | PDF

The second harmonic (SH) generated by nonlinear interaction in the RMR is a high-order TE2 mode. This can be coupled to the bus waveguide and converted into the fundamental TE0 mode using an asymmetric directional coupler (ADC). This reduces transmission loss of the SH mode and avoids unpredictable losses caused by directly coupling the high-order mode with the fiber at the chip's end face. In conventional ADCs, light can be coupled from one waveguide to another when the propagation constants of the eigenmodes in the two waveguides match. Considering the technological immaturity of lithium niobate etching, deviation in the waveguide width often arises from process instability. However, shortcuts to adiabaticity (STA) offer directional couplers with short device lengths and large fabrication tolerances [31,32]. The shape of the STA-enabled ADC at the RMR bottom is optimized using fast adiabatic evolution theory for efficient SH output generated via nonlinear conversion in the ring. The structure of the STA-ADC is depicted in Fig. 2(c), coupling the single-mode TE0 narrow waveguide and multi-mode TE2 wide waveguide. The electric field amplitudes Es1 and Es2 in the narrow and wide waveguides, respectively, evolve along the propagation length z according to the coupling mode equations:

$$\frac{d}{{dz}}\left[ \begin{array}{l} {E_{s1}}\\ E{}_{s2} \end{array} \right] ={-} i\left[ \begin{array}{l} - \Delta (z)\textrm{ }K(z)\\ K(z)\textrm{ }\Delta (z) \end{array} \right]\left[ \begin{array}{l} {E_{s1}}\\ E{}_{s2} \end{array} \right]$$
where K(z) represents the coupling coefficient and Δ(z) represents the mismatch of the propagation constant. By using the invariant STA method and setting boundary conditions of input coupling strength c(0) = 0 and output coupling strength c(Lz) = 0.06 based on requirements, the optimal shape of the ADC is obtained by solving Eq. (1). Considering the propensity for sidewall scattering loss in narrow waveguides during fabrication, we fixed the narrow waveguide width ws1 at 0.3 µm and the coupling length Lz at 100 µm. Furthermore, Eq. (1) is solved by introducing perturbation theory to consider the influence of waveguide width error and waveguide spacing error on the performance of the coupler. This enables obtaining a robust ADC design. The parametric equations satisfying the boundary conditions are constructed as follows:
$$\theta (z) = \frac{\pi }{2} - \frac{\pi }{2}\cos (\frac{{\pi z}}{L})$$
$$\phi (z) = \frac{\pi }{2} + {a_1}\sin (\frac{{\pi z}}{L}) + {a_2}\sin (\frac{{3\pi z}}{L})$$
where L is the coupling length, and ${a_1}$, ${a_2}$ are the parameters to be optimized. Based on program optimization calculations, when ${a_1} ={-} 0.75$ and ${a_2} = 0.03$, the perturbation caused by process error has the least influence on coupler performance. Figures 2(d) and 2(e) show the evolution of the wide waveguide width ws2 and coupling gap gs along the device length. We modeled the parameters of the ADC designed by the STA method in FDTD. The simulation results (Fig. 2(f)) show that the transmittance at 780 nm was 0.05, which is nearly consistent with the parameters initially set in the STA. Fortuitously, the large refractive index mismatch renders the pump TE0 mode nearly uncoupled in the ADC, preserving the pump quality factor in the microring. Two linear tapers transition the coupled waveguide to a nonlinear phase-matched width, forming a closed RMR.

3. Device fabrication and experimental test results

Figure  3(a) presents the metallographic microscope image and scanning electron microscope (SEM) image of the coupling region in the fabricated RMR. The double-resonance RMR was prepared on a 300 nm thick X-cut LNOI substrate (by NanoLN). A 100 nm Cr metal layer was first deposited on the lithium niobate surface by electron beam evaporation (EBE) to serve as a hard mask for lithium niobate etching. Electron beam lithography (EBL) then defined the pattern on electron beam resist spun over the Cr layer. This pattern was sequentially transferred to the Cr and lithium niobate layers via inductively coupled plasma reactive ion etching (ICP-RIE) using Cl2 and Ar gases, respectively, fully etching through the 300 nm lithium niobate layer. Finally, 3 µm of SiO2 was deposited by plasma enhanced chemical vapor deposition (PECVD) at 300°C as a cladding layer.

 figure: Fig. 3.

Fig. 3. (a) Metallographic microscope image of the racetrack microring resonator. Scanning electron microscope (SEM) images showing (b) the coupling region at the top and (c) the coupling region at the bottom.

Download Full Size | PDF

The experimental setup for linear and nonlinear characterization of the RMR is illustrated in Fig. 4(a). A C-band telecommunication source and a near infrared broadband source were utilized for linear characterization. A 1560 nm narrow linewidth laser (NKT Koheras Boostik E15) with 1 nm tuning range serves as the pump to generate second-harmonic (SH) light. The pump is amplified by an erbium-doped fiber amplifier (EDFA) and its polarization is controlled by a polarization controller (PC) to ensure TE polarization on-chip. The sample is mounted on a thermoelectric cooler (TEC) to achieve double resonance via heating. The pump and SH light are coupled on and off chip through lensed fibers (LF). On the chip, we created a wider waveguide with the same end coupler as the RMR, for which the transmission loss of the waveguide can be ignored. The coupling efficiency between the chip and the fiber was obtained by measuring the input and output optical power. The coupling losses measured in the experiment were 5.18 dB/facet at 1560 nm and 9 dB/facet at 780 nm. At the same time, we measured waveguides of different lengths and obtained losses of 1.5 dB/cm at 1560 nm and 2 dB/cm at 780 nm by measuring the transmission. After separation from residual pump light via a wavelength division multiplexer (WDM), the collected SH is spectrally analyzed by an optical spectrum analyzer (OSA). A 1560 nm photodetector (PD) monitors the pump power.

 figure: Fig. 4.

Fig. 4. (a) Experimental setup diagram. (b) Resonance spectra of the microring resonator at 28°C, with the black line representing the pump band and the red line representing the second-harmonic band. Transmission spectra under double resonance condition for (c) the 1560 nm pump light and (d) the 780 nm second-harmonic light.

Download Full Size | PDF

The RMR was first linearly characterized using a broadband light source covering the telecommunication and near infrared bands. The resonant wavelengths of the RMR can be expressed as ${\lambda _t} = {\lambda _{t0}} + {d_t} \cdot T$ and ${\lambda _v} = {\lambda _{v0}} + {d_v} \cdot T$ for the telecommunication and near infrared bands, respectively, where ${d_t} = 0.028nm/K$ and ${d_v} = 0.022nm/K$ denote the thermal-optic coefficients of the resonant spectra. Due to the disparity between ${d_t}$ and ${d_v}$, increasing the RMR temperature gradually reduces the wavelength detuning $\Delta \lambda = {\lambda _t} - 2{\lambda _v}$. At $\Delta \lambda = 0$, the two cavity resonant wavelengths satisfy the energy conservation requirement for ${\lambda _t} = 2{\lambda _v}$, achieving double resonance. As depicted in Fig. 4(b), setting the RMR temperature to 28°C aligns the resonant peaks at 1560.4 nm and 780.2 nm, fulfilling the double resonance condition. Figure 4(c) and 4(d) exhibit the resonant spectra under double resonance with loaded quality factors of ${Q_t} = 1 \times {10^5}$ and ${Q_v} = 2.8 \times {10^4}$ for the pump and second-harmonic, respectively.

The RMR was then nonlinearly characterized using a continuous 1560 nm narrow linewidth laser. Due to instabilities in the fabrication, some deviations exist between the fabricated and designed device dimensions. In principle, the SHG efficiency in the waveguide can be expressed as:

$$\eta \textrm{ = }\frac{{{P_{SH}}}}{{P_{pump}^2}} = {g^2}{L^2}\textrm{sin}{\textrm{c}^2}(\frac{{\Delta \beta L}}{2})$$
where g2 represents the nonlinear coupling strength related to the mode field overlap, effective mode area, and effective nonlinear coefficient. L denotes the waveguide length and Δβ is the phase mismatch. However, errors in the waveguide width directly induce nonlinear phase mismatch, preventing maximum SHG efficiency. As shown in Fig. 5(a), the experimental error was first calibrated by measuring SHG in graded-width straight waveguides to empirically obtain the optimal phase-matched width. The blue dots indicate measured SH powers, with the red curve showing a $\textrm{sin}{\textrm{c}^2}({{{\Delta \beta L} / 2}} )$ fit. Some data points exceed the fit curves, potentially stemming from nonuniform chip thickness and inconsistent waveguide etch widths.

 figure: Fig. 5.

Fig. 5. (a) SH power versus waveguide width for straight waveguides. (b) Comparison of SHG in single-resonant microring resonator at 28°C and double-resonant microring resonator at 45°C. (c) SHG contrast between direct waveguide and microring resonator at identical pump power. (d) On-chip SH power as a function of the squared pump power.

Download Full Size | PDF

SHG was then experimentally characterized in the optimally-width RMR to ensure maximum phase matching. At 28°C, the resonance spectrum fulfilled the double resonance condition for energy conservation, as shown in Fig. 5(b). The measured SH power was -36.4 dBm. Heating to 45°C detuned the double resonance by adjusting the pump wavelength to resonance, yielding -46.6 dBm SH power. Hence, double resonance provides a 10.2 dB enhancement in SHG efficiency compared to the single resonance state. The illustration in Fig. 5(b) depicts SHG in our RMR pumped at 1560 nm, with 780 nm near infrared light generated through second-order nonlinearity and captured by a CCD above the chip. A conspicuous red spot is observable at the chip's output facet coupled to the lensed fiber, shown at right. A 4 mm long straight waveguide with the same phase matching design was fabricated near the RMR to serve as a control. Figure 5(c) contrasts SHG in the straight waveguide and double-resonant RMR at the same pump power. The RMR affords a 25.73 dB increase in efficiency over the waveguide. Figure 5(d) shows the quadratic SH power dependence on pump power; the blue dots are experimental data and the red line denotes a quadratic fit. Under low pump power, the on chip SHG power ${P_{SH}}$ follows ${P_{SH}} \propto P_{pump}^2$, with a linear fit slope of 1.01 confirming the expected quadratic scaling. The normalized SHG efficiency is $\eta = \frac{{{P_{SH}}}}{{P_{pump}^2}} = 9972\%/W$. When the square of the pump optical power exceeds 0.3mW2, the SH power no longer increases. This saturation may be caused by the frequency-doubled light conversion efficiency reaching its limit. Another important reason is that the SH light power generated in the cavity becomes high enough to lead to the inverse process of SHG (that is, optical parametric oscillation) whereby the SH light gets converted back to the pump wavelength [33]. Additionally, high intra-cavity power can produce photorefractive effects that detune the resonator, destroying the phase matching and energy conservation conditions. The high achieved conversion efficiency stems from optimized pump and SH coupling, and double resonance. Further improving lithium niobate etching to reduce waveguide loss can improve the quality factor and enhance efficiency.

4. Conclusion

In conclusion, we have designed a racetrack microring resonator on X-cut lithium niobate on insulator for enhanced second-harmonic generation through modal phase matching. Independent coupling of the pump and second-harmonic enables near critical coupling for both. Optimizing the bending radius in the microring resonator averts detrimental mode hybridization effects on stable high-order mode second-harmonic transmission, further improving its quality factor. Fabrication on the lithium niobate on insulator chip was performed via electron beam lithography and dry etching. At 28°C, double resonance was attained in the microring, yielding a normalized second-harmonic generation efficiency of 9972%/W. Compared to a 4 mm straight waveguide with identical mode phase matching design in our experiment, the resonator demonstrates a 25.73 dB enhancement in SHG. Moreover, the X-cut configuration readily allows on-chip transverse electrodes for electro-optic control. Our work paves the path toward integrated generation of two-photon entanglement and complex quantum photonic circuits.

Funding

Wuhan National Laboratory for Optoelectronics; National Natural Science Foundation of China (61975062, U21A20511).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. D. Zhu, L. Shao, M. Yu, et al., “Integrated photonics on thin-film lithium niobate,” Adv. Opt. Photon. 13(2), 242–352 (2021). [CrossRef]  

2. Y. Qi and Y. Li, “Integrated lithium niobate photonics,” Nanophotonics 9(6), 1287–1320 (2020). [CrossRef]  

3. R. R. Xie, G. Q. Li, F. Chen, et al., “Microresonators in lithium niobate thin films,” Adv. Opt. Mater. 9(19), 2100539 (2021). [CrossRef]  

4. G. Poberaj, H. Hu, W. Sohler, et al., “Lithium niobate on insulator (LNOI) for micro-photonic devices,” Laser Photonics Rev. 6(4), 488–503 (2012). [CrossRef]  

5. Y. Li, Z. Huang, W. Qiu, et al., “Recent progress of second harmonic generation based on thin film lithium niobate,” Chin. Opt. Lett. 19(6), 060012 (2021). [CrossRef]  

6. C. Wang, X. Xiong, N. Andrade, et al., “Second harmonic generation in nano-structured thin-film lithium niobate waveguide,” Opt. Express 25(6), 6963–6973 (2017). [CrossRef]  

7. R. Luo, Y. He, H. Liang, et al., “Highly tunable efficient second-harmonic generation in a lithium niobate nanophotonic waveguide,” Optica 5(8), 1006–1011 (2018). [CrossRef]  

8. I. Briggs, S. Hou, C. Cui, et al., “Simultaneous type-I and type-II phase matching for second-order nonlinearity in integrated lithium niobate waveguide,” Opt. Express 29(16), 26183–26190 (2021). [CrossRef]  

9. R. Luo, H. Jiang, S. Rogers, et al., “On-chip second-harmonic generation and broadband parametric down-conversion in a lithium niobate microresonator,” Opt. Express 25(20), 24531–24539 (2017). [CrossRef]  

10. J. Y. Cheng, Y. M. Sua, H. Fan, et al., “Modal phase matched lithium niobate nanocircuits for integrated nonlinear photonics,” OSA Continuum 1(1), 229–242 (2018). [CrossRef]  

11. C. Yan, S. Wang, S. Zhao, et al., “Efficient and temperature-tunable second-harmonic generation in a thin film lithium niobate on insulator microdisk,” Appl. Phys. Lett. 122(10), 103504 (2023). [CrossRef]  

12. H. Jiang, H. Liang, R. Luo, et al., “Nonlinear frequency conversion in one dimensional lithium niobate photonic crystal nanocavities,” Appl. Phys. Lett. 113(2), 021104 (2018). [CrossRef]  

13. J. Lu, J. B. Surya, X. Liu, et al., “Periodically poled thin-film lithium niobate microring resonators with a second-harmonic generation efficiency of 250,000%/W,” Optica 6(12), 1455–1460 (2019). [CrossRef]  

14. H. Zhang, Q. Li, H. Zhu, et al., “Second harmonic generation by quasi-phase matching in a lithium niobate thin film,” Opt. Mater. Express 12(6), 2252–2259 (2022). [CrossRef]  

15. R. Luo, Y. He, H. Liang, et al., “Optical parametric generation in a lithium niobate microring with modal phase matching,” Phys. Rev. Appl. 11(3), 034026 (2019). [CrossRef]  

16. Z. Hao, L. Zhang, J. Wang, et al., “Sum-frequency generation of a laser and its background in an on-chip lithium-niobate microdisk,” Chin. Opt. Lett. 20(11), 111902 (2022). [CrossRef]  

17. J. J. Lu, M. Li, C. L. Zou, et al., “Toward 1% single-photon anharmonicity with periodically poled lithium niobate microring resonators,” Optica 7(12), 1654–1659 (2020). [CrossRef]  

18. A. Otterpohl, F. Sedlmeir, U. Vogl, et al., “Squeezed vacuum states from a whispering gallery mode resonator,” Optica 6(11), 1375–1380 (2019). [CrossRef]  

19. F. Mondain, T. Lunghi, A. Zavatta, et al., “Chip-based squeezing at a telecom wavelength,” Photonics Res. 7(7), A36–39 (2019). [CrossRef]  

20. J. J. Lu, A. A. Sayem, Z. Gong, et al., “Ultralow-threshold thin-film lithium niobate optical parametric oscillator,” Optica 8(4), 539–544 (2021). [CrossRef]  

21. C. Lu, Y. Zhang, J. Qiu, et al., “Highly tunable birefringent phase-matched second-harmonic generation in an angle-cut lithium niobate-on-insulator ridge waveguide,” Opt. Lett. 47(5), 1081–1084 (2022). [CrossRef]  

22. J. Lin, Y. Xu, J. Ni, et al., “Phase-matched second-harmonic generation in an on-chip LiNbO3 microresonator,” Phys. Rev. Appl. 6(1), 014002 (2016). [CrossRef]  

23. J. Lin, N. Yao, Z. Hao, et al., “Broadband quasi-phase-matched harmonic generation in an on-chip monocrystalline lithium niobate microdisk resonator,” Phys. Rev. Lett. 122(17), 173903 (2019). [CrossRef]  

24. J. Y. Chen, Z. H. Ma, Y. M. Sua, et al., “Ultra-efficient frequency conversion in quasi-phase-matched lithium niobate microring,” Optica 6(9), 1244–1245 (2019). [CrossRef]  

25. J. Y. Chen, C. Tang, Z. H. Ma, et al., “Efficient and highly tunable second-harmonic generation in Z-cut periodically poled lithium niobate nanowaveguid,” Opt. Lett. 45(13), 3789–3792 (2020). [CrossRef]  

26. C. Wang, C. Langrock, A. Marandi, et al., “Ultrahigh-efficiency wavelength conversion in nanophotonic periodically poled lithium niobate waveguide,” Optica 5(11), 1438–1441 (2018). [CrossRef]  

27. M. Xu, M. He, H. Zhang, et al., “High-performance coherent optical modulators based on thin-film lithium niobate platform,” Nat. Commun. 11(1), 3911 (2020). [CrossRef]  

28. D. E. Zelmon, D. L. Small, and D. Jundt, “Infrared corrected Sellmeier coefficients for congruently grown lithium niobate and 5 mol. % magnesium oxide-doped lithium niobate,” J. Opt. Soc. Am. B 14(12), 3319–3322 (1997). [CrossRef]  

29. I. H. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am. 55(10), 1205–1208 (1965). [CrossRef]  

30. D. X. Dai and M. Zhang, “Mode hybridization and conversion in silicon-on-insulator nanowires with angled sidewalls,” Opt. Express 23(25), 32452–32464 (2015). [CrossRef]  

31. D. Guo and T. Chu, “Silicon mode (de)multiplexers with parameters optimized using shortcuts to adiabaticity,” Opt. Express 25(8), 9160 (2017). [CrossRef]  

32. D. Guo and T. Chu, “Compact broadband silicon 3 dB coupler based on shortcuts to adiabatici,” Opt. Lett. 43(19), 4795–4798 (2018). [CrossRef]  

33. T. P. Mckenna, H. S. Stokowski, V. Ansari, et al., “Ultra-low-power second-order nonlinear optics on a chip,” Nat. Commun. 13(1), 4532 (2022). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) Schematic of double resonance RMR and the illustration shows the cross section of the LN waveguide. (b) Simulation of effective refractive index of pump and second harmonic light in straight waveguide. (c,d) Simulation of effective refractive index versus rotation angle in 180° curved waveguide. (e) FDTD simulation of TE2 and TM2 mode crosstalk for different bending radii. Insets show field distribution for radii of 26 µm and 30 µm.
Fig. 2.
Fig. 2. (a) Schematic diagram of the DC at the top of the RMR. (b)FDTD simulation results of the DC for 1560 nm pump and 780 nm SH. (c) Schematic diagram of the ADC at the bottom of the racetrack microring. The variation curves of ADC wide waveguide width ws2 (d) and coupling gap gz (e) with propagation length z. (f) FDTD simulation results of the ADC for 1560 nm pump and 780 nm SH.
Fig. 3.
Fig. 3. (a) Metallographic microscope image of the racetrack microring resonator. Scanning electron microscope (SEM) images showing (b) the coupling region at the top and (c) the coupling region at the bottom.
Fig. 4.
Fig. 4. (a) Experimental setup diagram. (b) Resonance spectra of the microring resonator at 28°C, with the black line representing the pump band and the red line representing the second-harmonic band. Transmission spectra under double resonance condition for (c) the 1560 nm pump light and (d) the 780 nm second-harmonic light.
Fig. 5.
Fig. 5. (a) SH power versus waveguide width for straight waveguides. (b) Comparison of SHG in single-resonant microring resonator at 28°C and double-resonant microring resonator at 45°C. (c) SHG contrast between direct waveguide and microring resonator at identical pump power. (d) On-chip SH power as a function of the squared pump power.

Equations (4)

Equations on this page are rendered with MathJax. Learn more.

d d z [ E s 1 E s 2 ] = i [ Δ ( z )   K ( z ) K ( z )   Δ ( z ) ] [ E s 1 E s 2 ]
θ ( z ) = π 2 π 2 cos ( π z L )
ϕ ( z ) = π 2 + a 1 sin ( π z L ) + a 2 sin ( 3 π z L )
η  =  P S H P p u m p 2 = g 2 L 2 sin c 2 ( Δ β L 2 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.